+ All Categories
Home > Documents > Analyzing Multistationarity in Chemical Reaction Networks using...

Analyzing Multistationarity in Chemical Reaction Networks using...

Date post: 01-Feb-2021
Category:
Upload: others
View: 3 times
Download: 0 times
Share this document with a friend
24
Analyzing Multistationarity in Chemical Reaction Networks using the Determinant Optimization Method Bryan elix a , Anne Shiu b , Zev Woodstock c, a Department of Mathematics, University of Texas at Austin, RLM 8.100, 2515 Speedway Stop C1200, Austin, Texas 78712-1202, USA b Department of Mathematics, Texas A&M University, Mailstop 3368, College Station, Texas 77843–3368, USA c Department of Mathematics and Statistics, James Madison University, Roop Hall 305, MSC 1911, Harrisonburg, Virginia 22807, USA Abstract Multistationary chemical reaction networks are of interest to scientists and mathematicians alike. While some criteria for multistationarity exist, obtain- ing explicit reaction rates and steady states that exhibit multistationarity for a given network—in order to check nondegeneracy or determine stability of the steady states, for instance—is nontrivial. Nonetheless, we accomplish this task for a certain family of sequestration networks. Additionally, our results allow us to prove the existence of nondegenerate steady states for some of these se- questration networks, thereby resolving a subcase of a conjecture of Joshi and Shiu. Our work relies on the determinant optimization method, developed by Craciun and Feinberg, for asserting that certain networks are multistationary. More precisely, we implement the construction of reaction rates and multiple steady states which appears in the proofs that underlie their method. Further- more, we describe in detail the steps of this construction so that other researchers can more easily obtain, as we did, multistationary rates and steady states. Keywords: Mass-action kinetics, chemical reaction networks, multistationarity, determinant optimization method, steady states, degeneracy 1. Introduction Although many dynamical systems arising in applications exhibit bistability, there is no complete characterization of such systems. Even for the subclass * Please send correspondence to ZW. Email addresses: [email protected] (Bryan elix), [email protected] (Anne Shiu), [email protected] (Zev Woodstock) Preprint submitted to Applied Mathematics and Computation April 21, 2016
Transcript
  • Analyzing Multistationarity in Chemical ReactionNetworks using the Determinant Optimization Method

    Bryan Félixa, Anne Shiub, Zev Woodstockc,

    aDepartment of Mathematics, University of Texas at Austin, RLM 8.100, 2515 SpeedwayStop C1200, Austin, Texas 78712-1202, USA

    bDepartment of Mathematics, Texas A&M University, Mailstop 3368, College Station,Texas 77843–3368, USA

    cDepartment of Mathematics and Statistics, James Madison University, Roop Hall 305,MSC 1911, Harrisonburg, Virginia 22807, USA

    Abstract

    Multistationary chemical reaction networks are of interest to scientists and

    mathematicians alike. While some criteria for multistationarity exist, obtain-

    ing explicit reaction rates and steady states that exhibit multistationarity for a

    given network—in order to check nondegeneracy or determine stability of the

    steady states, for instance—is nontrivial. Nonetheless, we accomplish this task

    for a certain family of sequestration networks. Additionally, our results allow

    us to prove the existence of nondegenerate steady states for some of these se-

    questration networks, thereby resolving a subcase of a conjecture of Joshi and

    Shiu. Our work relies on the determinant optimization method, developed by

    Craciun and Feinberg, for asserting that certain networks are multistationary.

    More precisely, we implement the construction of reaction rates and multiple

    steady states which appears in the proofs that underlie their method. Further-

    more, we describe in detail the steps of this construction so that other researchers

    can more easily obtain, as we did, multistationary rates and steady states.

    Keywords: Mass-action kinetics, chemical reaction networks,

    multistationarity, determinant optimization method, steady states, degeneracy

    1. Introduction

    Although many dynamical systems arising in applications exhibit bistability,

    there is no complete characterization of such systems. Even for the subclass

    ∗Please send correspondence to ZW.Email addresses: [email protected] (Bryan Félix), [email protected]

    (Anne Shiu), [email protected] (Zev Woodstock)

    Preprint submitted to Applied Mathematics and Computation April 21, 2016

  • of chemical kinetics systems and even under the assumption of mass-action

    kinetics, which is the focus of this work, the problem is difficult.5

    Here we consider the simpler, yet still challenging, question: which chemical

    reaction networks are multistationary, i.e. which have the capacity to exhibit

    two or more steady-state concentrations with the same reaction rates? Mathe-

    matically, this asks: among certain parametrized families of polynomial systems,

    which admit multiple positive roots? Therefore, this is a real algebraic geometry10

    problem, and we do not expect an easy answer in general.

    The first partial answers to this question are due to Feinberg, Horn, and

    Jackson in the 1970s. Their results in chemical reaction network theory [1, 2]

    (specifically, deficiency theory [3]) can preclude or guarantee multistationarity

    for certain classes of networks. For a survey of these and other methods, see [4].15

    Our work pertains to two related results: (1) a method for “lifting” multiple

    steady states from small networks to larger ones, and (2) the so-called determi-

    nant optimization method for certifying that a given network is multistationary.

    The lifting result, stated informally, is as follows: if a chemical reaction

    network contains an “embedded” network that is multistationary, then the entire20

    reaction network also is multistationary under certain hypotheses [5]. Therefore

    we are interested in cataloguing the multistationary networks which contain no

    embedded multistationary networks, because all larger multistationary networks

    contain at least one embedded multistationary subnetwork from the catalogue.

    As a step toward such a catalogue, Joshi and Shiu identified a certain infinite25

    family of chemical reaction networks K̃m,n to be of particular interest among all

    networks that include inflow and outflow reactions [4]. This family is minimal,

    in that it has no embedded subnetworks (with inflow and outflow reactions)

    that exhibit multistationarity. To analyze these networks, Joshi and Shiu used

    the second method for analyzing multistationarity mentioned above.30

    Developed by Craciun and Feinberg, the determinant optimization method

    can assert that a network is multistationary [6, 7]; as such, it is a partial converse

    to their results on “injective” reaction networks which guarantee that a network

    is not multistationary. This topic of injectivity has seen much interest in recent

    years (see [6, 8, 9] and the references therein); however, the determinant opti-35

    mization method has garnered comparatively little attention. The only related

    results that we are aware of are due to Banaji and Pantea [8], Feliu [10], and

    Müller et al. [9].

    Using the determinant optimization method, Joshi and Shiu proved that

    K̃m,n is multistationary for all integers m ≥ 2 and odd integers n ≥ 3 [4].40Furthermore, they conjectured that these networks can exhibit multiple nonde-

    2

  • generate steady states. (It is not guaranteed that the determinant optimization

    method produces nondegenerate steady states; we show this for the first time

    in Remark 3.11.) The significance of the conjecture is that if it is true, then

    K̃2,n would be the first example of an infinite family of chemical reaction net-45

    works with inflow and outflow reactions and at-most-bimolecular reactants and

    products—that is, minimal with respect to the embedding relation among all

    such networks which have the capacity to exhibit multiple nondegenerate steady

    states. Nondegeneracy is important because results that “lift” multiple steady

    states from embedded subnetworks or other typically smaller networks require50

    the steady states to be nondegenerate; a summary of such results appears in [4,

    §4]. Also, because trimolecular reactants/products are rather uncommon inchemistry and K̃2,n is at most bimolecular, this family of networks is of partic-

    ular interest in chemical applications.

    In the current work, we resolve the conjecture for the case n = 3 and all55

    m ≥ 2; in other words, we prove that K̃m,3 has the capacity to admit multiplenondegenerate steady states for all m ≥ 2 (Theorem 4.5). To accomplish this,we need information beyond the mere existence of multiple steady states; we also

    need precise values (or at least estimates) for the rates and steady states. By

    applying the proofs underlying the determinant optimization method in Craciun60

    and Feinberg’s work to the networks K̃m,n, we obtain (via standard methods for

    analyzing recurrence relations) explicit closed forms for multistationary rates

    and steady states. Then we use these closed forms to verify that the steady

    states are nondegenerate for small values of m and n.

    Finally, recognizing the usefulness of generating closed forms (or at least65

    estimates1) for rates and steady states for any reaction network that satisfies

    the hypotheses of the determinant optimization method, in Section 3 we outline

    the steps of the method with enough generality to be used in other contexts.

    These steps are present in Craciun and Feinberg’s work but are spread out

    over several proofs, so our contribution here is to reorganize the method into a70

    concise procedure.

    An outline of our work is as follows. Section 2 introduces chemical systems

    and the main conjecture. Section 3 describes the determinant optimization

    method in detail. We use this method to resolve some cases of the main conjec-

    ture in Sections 4 and 5. Finally, a discussion appears in Section 6.75

    Notation. We denote the positive real numbers by R+ := {x ∈ R | x > 0},

    1For general networks, the determinant optimization method need not yield closed forms

    for the rates and steady states, but one can nonetheless obtain estimates.

    3

  • and the standard inner product in Rn by 〈−,−〉. The ith entry of a vector x isdenoted xi.

    2. Background

    This section introduces chemical reaction networks, their corresponding mass-80

    action kinetics systems, and the main object of our paper: the sequestration

    network K̃m,n.

    2.1. Mass-action kinetics systems

    Definition 2.1. A chemical reaction network G = {S, C,R} consists of threefinite sets:85

    1. a set of species S = {X1, X2, . . . , Xs},2. a set C of complexes, which are non-negative integer linear combinations

    of the species, and

    3. a set R ⊆ C × C of reactions.

    Example 2.2. The following chemical reaction network:

    X1 + X2 → X3X2 → X1 + X4 ,

    is entirely defined by:90

    1. the set of species S = {X1, X2, X3, X4},2. the set of complexes C = {X1 + X2, X3, X2, X1 + X4}, and3. the set of reactions R = {(X1 + X2, X3), (X2, X1 + X4)}.

    Any reaction network G = {S, C,R} is contained in the fully open extensionnetwork G̃ obtained by including all inflow and outflow reactions:95

    G̃ :={S, C ∪ S ∪ {0} , R∪{Xi ↔ 0}Xi∈S

    }. (1)

    In other words, the fully open extension of any network is obtained by adding

    the reactions X → 0 (inflow) and 0→ X (outflow) for all X ∈ S.As all the reactions take place, the concentrations of each of the species

    will change. We make use of mass-action kinetics to define a system of ordi-

    nary differential equations that describes, for each species, how its concentration100

    changes as a function of time. This ODE system is described by the stoichio-

    metric matrix Γ and the reactant vector R(x), which is a vector-valued function

    of the vector of species concentrations x.

    4

  • Definition 2.3. Let G = {S, C,R} be a network, and let {y1 → y′1, y2 →y′2, . . . , y|R| → y′|R|} be an ordering of the reactions.105

    1. The reaction vector of the reaction yi → y′i is the vector y′i− yi, viewed inR|S|. Note that yi → y′i is a slight abuse of notation, used to denote thereaction yi · X → y′i · X where X is the vector of all species. Explicitly,the vectors yi and y

    ′i only contain species coefficients.

    2. The stoichiometric matrix of G is the |S|×|R| matrix Γ whose kth column110is the reaction vector of yk → y′k.

    3. The reactant vector R(x) is the vector of length |R| whose kth entry is the(monomial) product:

    rkx(yk)11 x

    (yk)22 · · ·x

    (yk)|S||S| ,

    where rk ∈ R+ is the reaction rate of the kth reaction.

    Definition 2.4. The mass-action kinetics system of a network G = {S, C,R}and a vector of reaction rates (rk) ∈ R|R|+ is defined by the following system ofordinary differential equations:115

    dx

    dt= Γ ·R(x) . (2)

    Example 2.5. For the following network:

    A + 2Br→ 2A ,

    Γ =

    [1

    −2

    ]and R(x) =

    (rxAx

    2B

    ), so the mass-action kinetics system (2) is:

    [dxAdt

    dxBdt

    ]= Γ ·R(x) =

    [rxAx

    2B

    −2rxAx2B

    ].

    An important characteristic of mass-action kinetics systems is that they may

    or may not have the capacity to admit (positive) steady states:

    Definition 2.6. A positive steady state is a vector x∗ ∈ R|S|>0 such that Γ ·R(x∗) = 0. A steady state x∗ is nondegenerate if Im(df(x∗)|Im(Γ)) = Im(Γ),where df(x∗) denotes the Jacobian matrix of the mass-action kinetics system at120

    x∗.

    5

  • Definition 2.7. A network that includes all inflow and outflow reactions is mul-

    tistationary2 if there exist two distinct concentration vectors x∗,x# and positive

    reaction rates such that Γ ·R(x∗) = Γ ·R(x#) = 0.

    2.2. The sequestration network Km,n125

    The main object of our paper is the fully open extension of the network

    Km,n:

    Definition 2.8. For positive integers n ≥ 2 and m ≥ 2, the sequestrationnetwork Km,n is:

    X1 + X2 → 0 (3)

    X2 + X3 → 0...

    Xn−1 + Xn → 0

    X1 → mXn .

    K̃m,n is the fully open extension of Km,n, obtained by adjoining all inflow and

    outflow reactions, as in (1).

    Schlosser and Feinberg analyzed variations of K̃2,n [11, Table 1], as did130

    Craciun and Feinberg [6, Table 1.1]. Joshi and Shiu introduced the version of

    the sequestration networks in Definition 2.8, and proved that some of them are

    multistationary:

    Proposition 2.9. [4, Lemma 6.9] For positive integers m ≥ 2 and n ≥ 3, if nis odd, then K̃m,n admits multiple positive steady states.135

    For m = 1 or n even, the network K̃m,n is “injective” and therefore not

    multistationary [4, §6]. Joshi and Shiu conjectured that Proposition 2.9 extendsas follows:

    Conjecture 2.10. For positive integers m ≥ 2 and n ≥ 3, if n is odd, thenK̃m,n admits multiple nondegenerate steady states.140

    To resolve Conjecture 2.10, we must show that Im(df(x∗)) = Im(Γ) for two

    distinct positive steady states x∗. We will see in (4) below that Γ is full rank,

    so we need only show that det(df(x∗)) 6= 0 for two positive steady states x∗.

    2The focus of this work is on certain networks K̃mn, that include all flow reactions. For net-

    works that do not include all flow reactions, the definition of multistationary must incorporate

    the conservation relations in the network, if any.

    6

  • Remark 2.11. As mentioned in the introduction, networks in which all reac-

    tants and products are at most bimolecular—that is, each complex has the form145

    0, X, X + Y , or 2X—are the norm in chemistry. This is the case for the

    networks K̃2,n, so that the nth internal reaction is X1 → 2Xn.

    We end this section by displaying the matrices that define the mass-action

    kinetics system (2) defined by K̃m,n. We order the reactions as follows: first, we

    enumerate the n internal (or true) reactions listed in (3) (so, the first reaction150

    is X1 + X2 → 0, and so on), next are the n outflow reactions (so, the (n + 1)streaction is X1 → 0, and so on), and then we have the n inflow reactions (so, the(2n + 1)st reaction is 0 → X1, and so on). We will refer to the sets of internal(true), outflow, and inflow reactions as RT , RO, and RI , respectively.

    The stoichiometric matrix for K̃m,n is:155

    Γ =

    −1 0 0 . . . 0 −1−1 −1 0 . . . 0 0

    0 −1 −1. . .

    ...... −In In

    0 0 −1. . .

    ......

    ......

    . . .. . . −1 0

    0 0 0 . . . −1 m

    , (4)

    where In is the n× n identity matrix. The reactant vector is:

    R(x) =

    r1x1x2

    r2x2x3...

    rn−1xn−1xn

    rnx1

    rn+1x1

    rn+2x2...

    r2nxn

    r2n+1...

    r3n

    ,

    where the ri ∈ R+ are the reaction rates and each xi ∈ R+ is the concentration

    7

  • of each species Xi. The mass-action ODEs (2.4) are:

    ẋ1 =− r1x1x2 − rnx1 − rn+1x1 + r2n+1ẋi =− ri−1xi−1xi − rixixi+1 − rn+ixi + r2n+i for 2 ≤ i ≤ n− 1

    ẋn =− rn−1xn−1xn + mrnx1 − r2nxn + r3n .

    Thus the Jacobian matrix, df(x), is the following (n× n)-matrix

    −r1x2 − rn − rn+1 −r1x1 0 . . . 0 0

    −r1x2 −r1x1 − r2x3 − rn+2 −r2x2 . . ....

    ...

    0 −r2x3 −r2x2 − r3x4 − rn+3. . . 0 0

    ... 0 −r3x4. . . −rn−2xn−2 0

    0...

    .... . . −rn−2xn−2 − rn−1xn − r2n−1 −rn−1xn−1

    mrn 0 0 . . . −rn−1xn −rn−1xn−1 − r2n

    .

    (5)

    3. Constructing multiple steady states via the determinant optimiza-

    tion method

    The determinant optimization method3 was developed by Craciun and Feinberg to160

    show that certain chemical reaction networks are multistationary [6]. More precisely,

    the method guarantees that some networks (such as those that satisfy the ‘Input’

    conditions below) are necessarily multistationary. For instance, Joshi and Shiu showed

    that the networks K̃m,n (for m ≥ 2 and odd n ≥ 3) satisfy the ‘Input’ conditions, andthus concluded these networks are multistationary (Proposition 2.9).165

    In fact, the determinant optimization method also applies to some networks that

    do not satisfy the ‘Input’ conditions. To determine if this is the case for a given

    network, one must check whether a certain optimization problem has a solution (see

    Remark 3.4). If so, then the method guarantees that the network is multistationary.

    However, in many applications, it is useful not only to know that a network is170

    multistationary but also to have explicit steady-state concentrations and reaction rates

    that are witnesses to multistationarity. For instance, here we would like to determine

    whether the steady states are degenerate, whereas in other settings one might like to

    perform stability analysis.

    Fortunately, the proofs in [6, §4] that underlie the determinant optimization method175are constructive, up to one use of the Intermediate Value Theorem, so one can generate

    or at least approximate steady states and rates. This section describes the step-by-

    step procedure to do this; following our steps is easier than (although equivalent to)

    “backtracking” through the proofs. That is, our contribution here is to re-package

    the determinant optimization method into a constructive algorithm. We will see that180

    3Related techniques for establishing multistationarity appear in work of Banaji and Pan-

    tea [8, §4], Feliu [10, §2], and Müller et al. [9, §3.2].

    8

  • for some networks, such as K̃m,n, the method constructs closed forms for the steady

    states and rates.

    Determinant optimization method (constructive version)

    Input: Any chemical reaction network G with n = |S| species that contains all ninflow reactions such that there exist n reactions y1 → y′1, y2 → y′2, . . . , yn → y′n185among the internal (true) and outflow reactions RT ∪RO of G for which

    (I) det(y1, y2, ..., yn) · det((y1 − y′1), (y2 − y′2), ..., (yn − y′n)) < 0, and(II) there exists a vector η̃ ∈ Rn+ such that Σni=1η̃i(yi − y′i) ∈ Rn+ = R

    |S|+ .

    Output: A certificate of multistationarity of G: (approximations of) a positive re-

    action rate vector (ry→y′) ∈ R|R|+ and two positive concentration vectors x∗ and x#190which are both steady states of the mass-action system defined by G and (ry→y′).

    Steps: Described below.

    With an eye toward resolving Conjecture 2.10, K̃m,n will be our ongoing example.

    Example 3.1. For K̃m,n (with m ≥ 2 and n ≥ 3 odd), hypothesis (II) is satisfied bythe vector η̃ = (1, 1, . . . , 1,m+ 1, 1) [4, Lemma 6.9]. Hypothesis (I) was proven in [4,195

    Lemma 6.7], where the n reactions are precisely the n internal (true) reactions (3).

    Conveniently, these are the reactions labeled yi → y′i of the sequestration network, for1 ≤ i ≤ n, so our notation for the first n reactions—as well as the use of n for thenumber of species—matches that of the determinant optimization method.

    The steps below involve a certain linear transformation Tη; specifically, for η ∈200RRT∪RO , the linear transformation Tη : R|S| → R|S| is defined by:

    Tη(δ) =∑

    y→y′∈RT∪RO

    ηy→y′(y · δ)(y − y′) . (6)

    Equivalently, the matrix representation of Tη is d(−f)(1, 1, . . . , 1) where the rates aregiven by ri = ηi. In other words, this matrix is the Jacobian matrix of the mass-

    action system (2.4) defined by the internal and outflow reaction rates η (and any

    choice of inflows: they do not appear in the Jacobian matrix) at the concentration205

    vector (1, 1, . . . , 1).

    Example 3.2. For K̃m,n, the matrix representation of Tη is:

    η1 + ηn + ηn+1 η1 0 · · · 0η1 η1 + η2 + ηn+2 η2 · · · 00 η2 η2 + η3 + ηn+3 η3 0... 0

    . . .. . .

    ...

    0... · · · ηn−2 + ηn−1 + η2n−1 ηn−1

    −mηn 0 · · · ηn−1 ηn−1 + η2n .

    (7)

    From the Jacobian matrix (5), it is clear that this matrix (7) equals d(−f)(1, 1, . . . , 1),where the reaction rates are given by ri = ηi.

    9

  • The first step is to construct a (strictly positive) vector η− ∈ R|RT∪RO|+ , indexed by210all internal (true) and outflow reactions, such that

    (I) det(Tη−) < 0, and

    (II)∑

    y→y′∈RT∪ROη−y→y′(y − y

    ′) ∈ R|S|+ .

    Craciun and Feinberg proved that these conditions (I) and (II) are satisfied by a vector

    η− of the following form:

    η−y→y′ =

    λη̃y→y′ if y → y′ ∈ {yi → y′i | i ∈ [n]}� else ,where λ is sufficiently large and � is sufficiently small [6, proof of Theorem 4.2].

    Example 3.3. For K̃m,n (with m ≥ 2 and n ≥ 3 odd), we define η− as follows:

    η−i =

    λ if 1 ≤ i ≤ n− 2 or i = n

    (m+ 1)λ if i = n− 1

    � if n+ 1 ≤ i ≤ 2n .

    Remark 3.4 (Stronger versions of the determinant optimization method). This sec-215

    tion describes the simplest version of the determinant optimization method. In fact,

    even if a network does not satisfy the hypotheses in the input given above, the method

    can still apply: [6, Remark 4.1] describes, in this setting, how to implement the above

    first step, i.e. how to test whether a suitable η− exists, as an optimization problem.

    Specifically, this is a polynomial optimization problem with linear constraints over a220

    compact set. Therefore, one can use any applicable optimization method. Additionally,

    see Remark 3.5 for how one can begin the algorithm at the second step.

    The second step is to construct a (strictly positive) vector η0 ∈ R|RT∪RO|+ for which:

    (I′) det(Tη0) = 0, and

    (II′)∑

    y→y′∈RT∪ROη0y→y′(y − y′) ∈ R

    |S|+ .225

    Craciun and Feinberg proved that this can be accomplished as follows [6, proof of

    Theorem 4.1]. First, construct an η+ ∈ R|RT∪RO|+ such that det(Tη+) > 0; do this byassigning a large value to outflow reactions and a small value to internal reactions:

    η+y→y′ =

    λ+ if y → y′ ∈ RO�+ if y → y′ ∈ RT ,where λ+ > 0 is large and �+ > 0 is small.

    Then, by interpolating between this vector η+ and the vector η− from the pre-

    vious step, the Intermediate Value Theorem (plus the fact that the set of vectors

    satisfying condition (II) is convex) guarantees the existence of an η0 with the required

    properties. Moreover, such a suitable η0 can be numerically approximated, and, with230

    careful tracking of error, one can use this approximation to generate steady states and

    concentrations in the following steps.

    10

  • Remark 3.5 (Interpretation of the second step and subsequent steps). What the

    second step does is to find reaction rates (given by η0 for the internal and outflow

    rates, and the vector in (II ′) for the inflow rates) at which the concentration vector235

    (1, 1, . . . , 1) is a degenerate steady state: degeneracy is by (I ′), and being a steady state

    comes from (II ′).

    Equivalently, if (r̃y→y′) is any positive vector of reaction rates at which some con-

    centration vector c̃ is a degenerate positive steady state, then the vector η ∈ R|RT∪RO|+defined coordinate-wise by ηy→y′ = r̃y→y′ c̃

    y satisfies the second step. So, a reader240

    who has already found a degenerate positive steady state of their system could start the

    determinant optimization method at the second step. In other words, one could begin

    applying the method by immediately searching for a suitable η0 (without first generating

    η− and η+). One strategy for doing this is described in Remark 3.6, which we employ

    for K̃m,n beginning in Example 3.7.245

    In the next steps, the determinant optimization method constructs a certain vector

    δ so that |δ| is a suitable bifurcation parameter: for |δ| small but positive, the degeneratesteady state breaks into two nondegenerate steady states.

    Remark 3.6. Here is one strategy for constructing a suitable η0 (without using η−

    and η+). First, identify (if possible) an η ∈ R|S|+ and a reaction yi → y′i among the250internal (true) and outflow reactions such that:

    (a)∑

    y→y′∈(RT∪RO)\{yi→y′i}η0y→y′(y − y′) ∈ R

    |S|+ , and

    (b) yi − y′i ∈ R|S|≥0 (this holds, for instance, if yi → y

    ′i is an outflow reaction).

    One could see whether η+ or η− might work (we use η− in Example 3.7 below). Then,

    define η as follows: let the entry η0i (corresponding to the same ith reaction) be free, and255

    fix η0j = η−j for j 6= i. Then, solve the (univariate polynomial) equation det(Tη0) = 0.

    If there is a positive solution (for η0i ), then the resulting vector η0 is positive, and (I ′)

    holds by construction. Furthermore, (II ′) holds because the sum in (II ′) is precisely the

    sum of a positive vector (namely, the sum in (a)) and a non-negative vector (namely,

    η0i (yi − y′i)). However, η0i is not guaranteed to be positive, so this strategy may fail.260

    Example 3.7. For K̃m,n (with m ≥ 2 and n = 3, 5, 7, 9, 11), the following choice ofη0 satisfies the requirements of the second step:

    η0i =

    λ if 1 ≤ i ≤ n− 2 or i = n

    (m+ 1)λ if i = n− 1

    � if n+ 1 ≤ i ≤ 2n− 1(m+1)(mλn+λ2(m+1)kn−2)(λ(m+2)+�)kn−2−λ2kn−3

    − λ(m+ 1) if i = 2n ,

    (8)

    where λ > 0 and � > 0 are such that η02n is positive4 (thus, all coordinates of η0 are

    4We checked that such λ, � exist for n = 3, 5, 7, 9, 11 (and m ≥ 2), and we furthermoreconjecture that for larger n, choosing λ sufficiently large and � sufficiently small will suffice.

    11

  • positive), and ki is ith principal minor of the matrix representation of Tη0 displayedbelow in (9), i.e. ki is the determinant of the i× i (tridiagonal) upper-left submatrix265of (9) (also, k0 := 1).

    To show that this choice of η0 satisfies the two conditions of the second step, we

    first note that (II ′) is straightforward to verify.

    Satisfying (I ′) only requires η0 to satisfy one (determinantal) equation, so allowing

    one free variable is sufficient. We choose η02n as this free variable, and we will recover270

    the formula in (8). Namely, we let all other coordinates η0i have the form given above

    (for 1 ≤ i ≤ 2n− 1), and then, recalling (7), the matrix representation of Tη0 is:

    2λ+ � λ 0 · · · 0λ 2λ+ � λ 0 · · · 0

    0 λ. . .

    . . ....

    ... 0. . . 0

    0... λ+ λ(m+ 1) + � λ(m+ 1)

    −mλ 0 · · · 0 λ(m+ 1) λ(m+ 1) + η02n

    . (9)

    Expanding (9) along the bottom row, we obtain the determinant of Tη0 :

    det(Tη0) = (−1)nm(m+ 1)λn − λ2(m+ 1)2kn−2 + (λ(m+ 1) + η02n)kn−1 , (10)

    where we recall that ki is ith principal minor of Tη0 .The determinant of tridiagonal matrices can be solved recursively [12]. For our

    matrix, it is easy to verify the following recursion for i ≤ n− 2, which is independentof m:

    ki+2 = (2λ+ �)ki+1 − λ2ki ,

    with initial values k0 := 1 and k1 = 2λ+�. Notice that kn−1 must be treated separately,because the (n − 1)st row contains η0n−1, which is a function of m. Using standardmethods we can get the generating function of the recurrence.

    ki =1

    2i+1c1∗(−c2(c2 − c1)i + c1(c2 − c1)i + c2(c1 + c2)i + c1(c1 + c2)i

    ),

    where c1 = (�)12 (� + 4λ)

    12 and c2 = � + 2λ. Notice here that ki is always positive

    for sufficiently small �. A formula for kn−1 is given using the formula for tridiagonalmatrices [12]:

    kn−1 = kn−2(λ(m+ 2) + �)− λ2kn−3 .

    With this recurrence solved, we set equation (10) to zero and then derive the explicit275

    function for η02n from (10) in terms of m,λ, and �; this is the formula in (8).

    The third step is to construct a nonzero vector δ ∈ R|S| in the nullspace of Tη0 , i.e.such that Tη0 · δ = 0. (Such a vector δ exists because det(Tη0) = 0.)

    12

  • Example 3.8. For our example K̃m,n (with m ≥ 2 and n ≥ 3 odd), we claim that thevector δ whose coordinates are defined as follows is in the nullspace of (9):280

    δk =

    δ1 if k = 1

    −(2λ+�)λ

    δk−1 − δk−2 if 2 ≤ k ≤ n− 1−(λ(m+2)+�)

    λ(m+1)δn−1 − 1m+1δn−2 if k = n ,

    (11)

    where we introduce δ0 = 0 for convenience in solving the recurrence relation, and

    δ1 6= 0 is our free variable. Note that the last coordinate, δn, has a different formula,because the (n− 1)st row in (9) used to define δn contains terms dependent on m thatdo not satisfy the recurrence.

    To see that δ is a nonzero vector in the nullspace of Tη0 , notice that the conditions285

    on δ that state that its inner product with each of the the first n−1 rows of Tη0 coincideprecisely with the n− 1 recurrences in the definition of δ (11). We claim that the lastrow of Tη is linearly dependent on the other rows, and from this we will conclude that

    the last row of Tη0 automatically has zero inner product with δ. To see this, we recall

    that detTη0 = 0 by construction, so we need only show that the first n− 1 rows of Tη0290are linearly independent. Assume, to the contrary, that there is a non-trivial linear

    combination of the first n − 1st rows of Tη0 that adds to 0. Note that in the firstn− 1 rows only the last one contains an entry in the last column, namely η0n−1. Sinceη0n−1 6= 0, the corresponding scalar of the n−1st row should be 0, thus annihilating theentire row. In the same manner, the corresponding scalar for the n−2nd row would be295equal to 0. Repeating the process shows that the only linear combination that adds to

    0 is the trivial one, thus, arriving at a contradiction. So, δ is in the nullspace of Tη0 .

    Again, by using standard techniques for analyzing recurrences, we find the gener-

    ating function for each of the first n− 1 entries of δ:

    δk = δ1λ ·(√

    4λ�+ �2 − (2λ+ �))k − (−√

    4λ�+ �2 − (2λ+ �))k

    2kλk√

    4λ�+ �2,

    for 1 ≤ k ≤ n− 1.

    The final step is to use the vectors η0 and δ (or, as we will see, a sufficiently scaled

    version of δ) from the previous two steps to construct a certificate of multistation-

    arity [6, proof of Lemma 4.1]; namely, the internal (true) and outflow reaction rates

    are:

    ry→y′ =〈y, δ〉

    e〈y,δ〉 − 1η0y→y′ for all y → y′ ∈ RT ∪RO ,

    the inflow reaction rates are the coordinates of the following vector:

    (r0→Xi) =∑y→y′

    η0y→y′(y − y′) ∈ R|S|+ , (12)

    and the two steady states are:

    x∗ = (1, 1, ..., 1) and x# = (eδ1 , eδ2 , ..., eδ|S|) .

    13

  • Craciun and Feinberg showed that for sufficiently small scaling of δ, all inflow rates (12)300

    are positive [6].

    Example 3.9. For K̃m,n, with m ≥ 2 and n = 3, 5, 7, 9, 11, we checked that δ1 = 1suffices. Details for the n = 3 case are provided in Remark 4.2.

    Summarizing what we accomplished above, we have closed-form expressions for

    reaction rate constants and steady states that show that the sequestration network is305

    multistationary:

    Theorem 3.10. Consider positive integers m ≥ 2 and n ∈ {3, 5, 9, 11}. Let δ ∈ Rn

    be as in (11) with δ1 = 1. Also, let η0 be as is (8)5. Then, for the following internal

    (true) and outflow reaction rates:

    ri =〈yi, δ〉

    e〈yi,δ〉 − 1η0i for all i ∈ {1, 2, ..., 2n} ,

    and the following inflow reaction rates:

    r2n+1 = r1 + rn + rn+1

    r2n+i = ri−1 + ri + rn+i for all 2 ≤ i ≤ n− 1

    r3n = rn−1 + r2n −mrn ,

    the concentrations:

    x∗ = (1, 1, ..., 1) and x# = (eδ1 , eδ2 , ..., eδn) (13)

    both are positive steady states of the mass-action kinetics system defined by K̃m,n and

    the reaction rates ri above.

    Remark 3.11. One may wonder whether or not the determinant optimization method

    has the potential to create degenerate steady states. Indeed, if we could prove that310

    this method always constructs nondegenerate steady states, then this would resolve

    Conjecture 2.10. However, this is not the case.

    We determined this by analyzing K̃2,3 as in Theorem 3.10. By letting � be a free

    variable we compute the parametrized determinants det(df(x∗)) and det(df(x#)) as

    functions of �. Both functions are easily checked to be continuous for positive values315

    of �, and from the graph (Figure 1), we can see easily that there exist choices of �

    for which one of the two steady states is degenerate. More precisely det(df(x∗)) = 0

    for some � ∈ (0.12, 0.125) and det(df(x#)) = 0 for some � ∈ (0.240, 0.241) and some� ∈ (1.159, 1.160).

    5In fact, this theorem will hold for any larger n for which the last coordinate of η0 as

    defined in (8) can be made to be positive.

    14

  • Figure 1: Graphs of det(df(x∗)) solid and det(df(x#)) dashed as functions of �

    .

    4. Resolving Conjecture 2.10 for the n = 3 case320

    Recall that Conjecture 2.10 asserts that K̃m,n admits multiple nondegenerate pos-

    itive steady states, for integers m ≥ 2 and n ≥ 3 with n odd. The main result of thissection (Theorem 4.5) resolves the conjecture when n = 3. To accomplish this, we first

    write down rate constants for this n = 3 case for which there are two steady states

    x∗ and x#; these values were obtained by the determinant optimization method in325

    the previous section for the general K̃m,n case (Proposition 4.1). We then resolve the

    conjecture for n = 3 by proving that x∗ and x# are nondegenerate.

    4.1. Reaction rate constants for which K̃m,3 is multistationary

    Proposition 4.1 below specializes Theorem 3.10 to the n = 3 case. Following the

    description in Section 3, λ = 1 and � = 0.1 will suffice, and then we obtain

    η0 =

    (λ, λ(m+ 1), λ, �, �,

    m2 − 0.31m− 1.312.1m+ 3.41

    )Tfrom the second step of the determinant optimization method. Next, in the third step,

    we find that the following vector spans the nullspace of Tη0 :

    δ =

    (1, − 2.1, 2.1m+ 3.41

    m+ 1

    )T.

    Thus, Theorem 3.10 specializes to:

    15

  • Proposition 4.1. Consider any integer m ≥ 2, and the following internal, outflow,330and internal reaction rates:

    r1 =−1.1

    e−1.1−1 ≈ 1.65 r2 =1.31

    e1.31m+1−1

    r3 =1e−1 ≈ .58

    r4 =.1e−1 ≈ .06 r5 =

    −.21e−2.1−1 ≈ .24 r6 =

    m−1.31

    e2.1m+3.41

    m+1 −1r7 = r1 + r3 + r4 ≈ 2.29 r8 = r1 + r2 + r5 r9 = r2 + r6 −mr3 .

    (14)

    Then for the mass-action kinetics system defined by the fully open sequestration net-

    work K̃m,3 and the above rate constants ri, both x∗ = (1, 1, 1) and x# =

    (e, e−2.1, e

    2.1m+3.41m+1

    )are positive steady states.

    Note that in Proposition 4.1, only x#3 , r2, r6, r8, and r9 depend on m.335

    Remark 4.2. The only reaction rate in (14) that is not obviously positive is the inflow

    rate r9, so we verify it here:

    r9 = r2 + r6 −mr3 > r2 + 0−m(

    1

    e− 1

    )≥ m−m

    (1

    e− 1

    )> 0 ,

    where the second-to-last inequality follows from Lemma 4.3 below.

    4.2. Bounding rates and steady states of K̃m,3

    Here we give upper and lower bounds which we will use to prove that x∗ and x#

    are nondegenerate. The following bounds are on the third coordinate of x#:

    e2.1y+3.41

    y+1 ≥ x#3 = e2.1m+3.41

    m+1 > e2.1 for all m ≥ y ≥ 0 . (15)

    The first inequality in (15) follows from the easy fact that e2.1m+3.41

    m+1 is a decreasing340

    function when m > 0, and the second inequality is straightforward.

    The proofs of the following two upper/lower bounds are in Appendix A:

    Lemma 4.3 (Bounds on r2). When λ = 1 and � = 0.1, the rate constant r2 defined

    in (14) satisfies the following inequalities for all m ≥ 2:

    m+ 1 > r2 ≥ m .

    Lemma 4.4 (Bounds on r6). When λ = 1 and � = 0.1, the rate constant r6 defined

    in (14) satisfies the following inequalities:

    0.14m > r6 > 0.13m− 0.5 ,

    where the upper bound holds for m ≥ 2, and the lower bound holds for m ≥ 20.

    16

  • 4.3. Proving nondegeneracy of steady states for the network K̃m,3

    The main result of this section is:345

    Theorem 4.5 (Resolution of Conjecture 2.10 when n = 3). For integers m ≥ 2, thenetwork K̃m,3 has the capacity to admit multiple nondegenerate positive steady states.

    We will prove Theorem 4.5 by showing that x∗ and x# in Proposition 4.1 are

    nondegenerate, i.e. we must prove that the image of the 3× 3 Jacobian matrix df(x)at each of the steady states is equal to the image of the 3 × 9 matrix Γ. As stated350earlier (after Conjecture 2.10), since Γ is full rank, our problem reduces to showing

    that det(df(x)) 6= 0 for both steady states and for all integers m ≥ 2.We begin by displaying the Jacobian matrix (5) of K̃m,3:

    df(x) =

    −r1x2 − r3 − r4 −r1x1 0−r1x2 −r1x1 − r2x3 − r5 −r2x2mr3 −r2x3 −r2x2 − r6

    .Thus, our goal is to show that the following determinants (obtained by strategically

    cancelling and rearranging terms) are nonzero for all integers m ≥ 2:

    D1 := det(df(x∗)) = r2r1r3m − (r2 + r6)(r1r3 + r1r4 + r1r5 + r3r5 + r4r5)

    − r2r6(r1 + r3 + r4) (16)

    D2 := det(df(x#)) = r2x

    #2 ((r1x

    #2 + r3 + r4)(r2x

    #3 ) + r1x

    #1 mr3)

    − (r2x#2 + r6)(r1x#2 + r3 + r4)(r1x

    #1 + r2x

    #3 + r5)

    + (r2x#2 + r6)(r1x

    #1 r1x

    #2 ) (17)

    From its graph6, D1 appears to be increasing quadratically as a function of m. So,

    to prove that D1 is nonzero for integer values of m ≥ 2, we will bound it from below bya quadratic function. Similarly, D2 appears to be decreasing quadratically, so we will355

    bound it from above by another quadratic function. Using these bounds, we will then

    conclude that D1 and D2 are strictly positive and negative (respectively) after certain

    cutoff points of m, effectively showing nondegeneracy of both steady states beyond the

    cutoffs. Finally, we will complete the proof by evaluating D1 and D2 at the remaining

    integers m between the 2 and the cutoff points to show that these values are nonzero.360

    Proof of Theorem 4.5. We generate our rates and concentrations as in Proposition 4.1.

    Following the description immediately after Theorem 4.5, we need only show that

    D1 6= 0 and D2 6= 0 for all integers m ≥ 2. First, we bound D1 by using its formula (16)together with the bounds in Lemmas 4.3 and 4.4:

    D1 > m2r1r3 − ((m+ 1) + 0.14m)(r1r3 + r1r4 + r1r5 + r3r5 + r4r5)

    − (m+ 1)(0.14m)(r1 + r3 + r4) .

    6Analogous graphs for larger n appear in Appendix B.

    17

  • Next, estimating the remaining rates ri, which are constants (recall equations (14)),

    by appropriate upper or lower bounds, we obtain:

    D1 > m2(0.95)− ((m+ 1) + 0.14m)1.61− (m+ 1)(0.14m)2.29

    = 0.6294m2 − 2.156m− 1.61 .

    It is easy to show that the quadratic function which bounds D1 above is always positive

    for integers m > 4. So, D1 > 0 for m > 4. Thus, it remains only to show that D1 6= 0at m = 2, 3, 4; indeed, those values are nonzero and are listed in Table 1.

    m 2 3 4 5 6 7 8 9 10

    D1 = det(df(x∗)) 0.336 2.784 6.525

    D2 = det(df(x#)) -1.063 -3.811 -7.85 -13.19 -19.8 -27.71 -36.89 -47.36 -59.11

    m 11 12 13 14 15 16 17 18 19

    D2 = det(df(x#)) -72.14 -86.4 -102 -118.9 -137.1 -156.5 -177.2 -199.2 -222.5

    Table 1: Determinants of the Jacobian matrices (16–17) at the two steady states x∗ and x#

    for the values of m ≥ 2 before the proven bounds are valid. All of these determinants arenonzero, so the corresponding steady states are nondegenerate.

    Now we proceed to bound D2. Again, we use its formula (17) together with the365

    bounds in (15) and Lemmas 4.3 and 4.4:

    D2 < (m+ 1)x#2 ((r1x

    #2 + r3 + r4)((m+ 1)x

    #3 ) + r1x

    #1 mr3)

    − (mx2 + (.13m− .5))(r1x#2 + r3 + r4)(r1x#1 +mx

    #3 + r5)

    + ((m+ 1)x#2 + .14m)(r1x#1 r1x

    #2 ) .

    Note that we used the lower bound on r6, so the above inequality holds for m ≥ 20(and thus we will need to check the values of m between 2 and 19 separately). In the

    same manner as before, we approximate all of the constants appropriately for m ≥ 20,and then simplify:

    D2 < (m+ 1).13((.85)((m+ 1)8.7) + 2.61m)− (.25m− .5)(.84)(4.72 + 8.16m)

    + (.13(m+ 1) + .14m)(.91)

    = −0.41295m2 + 4.9437m+ 3.06205 .

    Therefore, it is easy to see that D2 is nonzero for m ≥ 20. For 2 ≤ m ≤ 19 we againrefer to Table 1, which completes our proof.

    5. Resolving Conjecture 2.10 for small values of m and n370

    The main result of this section extends Theorem 4.5 to n ≤ 11, for small m:

    18

  • Theorem 5.1 (Resolution of Conjecture 2.10 for small m and n). For m = 2, 3, 4, 5

    and n = 5, 7, 9, 11, the network K̃m,n has the capacity to admit multiple nondegenerate

    positive steady states.

    Proof. We generate our rates and concentrations as in Theorem 3.10: it is straightfor-375

    ward to check that δ1 = 1, λ = 1, and � = 0.001 satisfy all necessary hypotheses. Thus,

    we obtain two steady states, x∗ = (1, 1, . . . , 1) and x# defined in (13). Then, as in the

    proof of Theorem 3.10, we verify that the determinants det(df(x∗)) and det(df(x#))

    are nonzero for m = 2, 3, 4, 5, which is readily seen from their graphs, which appear in

    Appendix B. (In fact, the graphs strongly suggest that the conjecture holds completely380

    for each of these values of n, namely n = 5, 7, 9, 11, i.e. for m > 5 as well.)

    6. Discussion

    As stated in the introduction, deciding whether a chemical reaction network is

    multistationary is not easy in the general case. And even when we can confirm that

    a network is multistationary, there is no general technique to show that it will admit385

    multiple nondegenerate steady states. Nonetheless, in this paper we succeeded in this

    task for certain sequestration networks K̃m,n by using the determininant optimization

    method to obtain closed forms for reaction rates and steady states.

    Our work resolved the n = 3 case of Conjecture 2.10, and we believe that our

    results form an important step toward resolving the full conjecture. Specifically, one390

    could use the formulas for rates and steady states given in Theorem 3.10 to analyze

    the general case, or, perhaps easier, the case of some fixed m and general n. Two

    other possible approaches are to (1) find an alternate method to obtain closed forms

    for the steady states of a chemical reaction network, or (2) identify criteria that can

    guarantee that steady states are nondegenerate.395

    Expanding on the last idea, our ultimate goal is to develop general techniques

    to assert that steady states of a chemical reaction network are nondegenerate. For

    instance, our analysis of the Jacobian determinants in this work suggest that even

    if the determinant optimization method yields a degenerate steady state, then the

    rate constants can be perturbed slightly so that the degenerate steady state becomes400

    nondegenerate (and the other steady state also remains nondegenerate). Is this true

    for any network for which the determinant optimization method applies? If so, then

    this would completely resolve Conjecture 2.10, and, more generally, this would enable

    us to more readily “lift” multistationarity and thereby enlarge our catalogue of known

    multistationary networks.405

    Acknowledgements

    BF and ZW conducted this research as part of the NSF-funded REU in the Department

    of Mathematics at Texas A&M University (DMS-1460766), in which AS served as

    mentor. All authors contributed substantially to this work. AS was supported by

    19

  • the NSF (DMS-1312473). The authors thank Dean Baskin for help with the proof410

    of Lemma 4.3, and Maya Johnson, Badal Joshi, Emma Owusu Kwaakwah, Casian

    Pantea, Xiaoxian Tang, and Jacob White for advice and fruitful discussions. The

    authors also thank an anonymous referee.

    References

    [1] M. Feinberg, Chemical oscillations, multiple equilibria, and reaction network415

    structure, in: W. Stewart, W. Rey, C. Conley (Eds.), Dynamics and Modelling

    of Reactive Systems, Academic Press, 1980, pp. 59–130.

    [2] F. Horn, R. Jackson, General mass action kinetics, Arch. Ration. Mech. Anal.

    47 (2) (1972) 81–116.

    [3] M. Feinberg, Chemical reaction network structure and the stability of complex420

    isothermal reactors I. The deficiency zero and deficiency one theorems, Chem.

    Eng. Sci. 42 (10) (1987) 2229–2268.

    [4] B. Joshi, A. Shiu, A survey of methods for deciding whether a reaction network

    is multistationary, to appear in Mathematical Modelling of Natural Phenomena,

    special issue on Chemical Dynamics. Available from arXiv:1412.5257 (2015).425

    [5] B. Joshi, A. Shiu, Atoms of multistationarity in chemical reaction networks, Jour-

    nal of Mathematical Chemistry 51 (2012) 153–178.

    [6] G. Craciun, M. Feinberg, Multiple equilibria in complex chemical reaction net-

    works: I. the injectivity property, SIAM Journal of Applied Mathematics 65

    (2005) 1526–1546. doi:10.1137/S0036139904440278.430

    [7] G. Craciun, C. Pantea, Computational methods for analyzing bistability in bio-

    chemical reaction networks, Proceedings of 2010 IEEE International Symposium

    on Circuits and Systems (ISCAS) (2010) 549–552.

    [8] M. Banaji, C. Pantea, Some results on injectivity and multistationarity in chem-

    ical reaction networks, Available online at arXiv:1309.6771 (2013).435

    [9] S. Müller, E. Feliu, G. Regensburger, C. Conradi, A. Shiu, A. Dickenstein, Sign

    conditions for injectivity of generalized polynomial maps with applications to

    chemical reaction networks and real algebraic geometry, To appear in Foundations

    of Computational Mathematics. doi:10.1007/s10208-014-9239-3.

    [10] E. Feliu, Injectivity, multiple zeros and multistationarity in reaction networks,440

    Proceedings of the Royal Society of London A: Mathematical, Physical and En-

    gineering Sciences 471 (2173) (2014) 18 pages. doi:10.1098/rspa.2014.0530.

    20

    http://dx.doi.org/10.1137/S0036139904440278arXiv:1309.6771http://dx.doi.org/10.1007/s10208-014-9239-3http://dx.doi.org/10.1098/rspa.2014.0530

  • [11] P. M. Schlosser, M. Feinberg, A theory of multiple steady states in isothermal ho-

    mogeneous CFSTRs with many reactions, Chemical Engineering Science 49 (11)

    (1994) 1749–1767. doi:DOI:10.1016/0009-2509(94)80061-8.445

    URL http://www.sciencedirect.com/science/article/B6TFK-44CRH1D-5C/2/

    4368d27af5ad95131107f482727432a8

    [12] N. Cahill, D. Narayan, Fibonacci and Lucas numbers as tridiagonal matrix

    determinants, Fibonacci Quarterly 42N3 (2004) 216–221.

    URL http://scholarworks.rit.edu/cgi/viewcontent.cgi?article=2122&450

    context=article

    [13] Desmos, Inc., Desmos graphing calculator, http://www.desmos.com (2015).

    Appendix A: proofs of Lemmas 4.3 and 4.4.

    Lemma 6.1 (Lemma 4.3). The function r2(m) =1.31

    e1.31m+1−1

    satisfies the following

    inequalities for all m ≥ 2:

    m+ 1 > r2(m) ≥ m .

    Proof. For the upper bound, first observe that

    log

    ((1 +

    1.31

    m+ 1

    )m+1)< log(e1.31) = 1.31 for all m ≥ 0 ,

    since limx→∞(1 +yx

    )x converges to ey from below for positive values of y. Thus:

    1.31 > log

    ((1 +

    1.31

    m+ 1

    )m+1)

    > log

    ((m+ 2.31

    m+ 1

    )m+1)= (m+ 1) log

    (m+ 2.31

    m+ 1

    ),

    which implies that

    1.31

    m+ 1> log

    (m+ 2.31

    m+ 1

    )=⇒ e

    1.31m+1 >

    m+ 2.31

    m+ 1

    =⇒ e1.31m+1 (m+ 1)− (m+ 1) > 1.31 .

    and this final inequality implies our desired upper bound: m+1 > 1.31

    e1.31m+1−1

    = r2(m).

    Notice that our lower bound is equivalent to the following inequality:

    m+ 1.31

    m≥ e

    1.31m+1 for all m ≥ 2 . (18)

    21

    http://www.sciencedirect.com/science/article/B6TFK-44CRH1D-5C/2/4368d27af5ad95131107f482727432a8http://www.sciencedirect.com/science/article/B6TFK-44CRH1D-5C/2/4368d27af5ad95131107f482727432a8http://www.sciencedirect.com/science/article/B6TFK-44CRH1D-5C/2/4368d27af5ad95131107f482727432a8http://dx.doi.org/DOI: 10.1016/0009-2509(94)80061-8http://www.sciencedirect.com/science/article/B6TFK-44CRH1D-5C/2/4368d27af5ad95131107f482727432a8http://www.sciencedirect.com/science/article/B6TFK-44CRH1D-5C/2/4368d27af5ad95131107f482727432a8http://www.sciencedirect.com/science/article/B6TFK-44CRH1D-5C/2/4368d27af5ad95131107f482727432a8http://scholarworks.rit.edu/cgi/viewcontent.cgi?article=2122&context=articlehttp://scholarworks.rit.edu/cgi/viewcontent.cgi?article=2122&context=articlehttp://scholarworks.rit.edu/cgi/viewcontent.cgi?article=2122&context=articlehttp://scholarworks.rit.edu/cgi/viewcontent.cgi?article=2122&context=articlehttp://scholarworks.rit.edu/cgi/viewcontent.cgi?article=2122&context=articlehttp://scholarworks.rit.edu/cgi/viewcontent.cgi?article=2122&context=articlehttp://www.desmos.com

  • We set a = 1.31, make use of the change of variables z = 1m

    , and then apply log to

    see that our desired inequality (18) is equivalent to:

    log(1 + az) ≥ az−1 + 1

    =az

    1 + zfor all z ∈ (0, 1/2) .

    We will show that log(1 + az)− az1+z≥ 0. To this end, define b by 1− b = a− 1, and

    notice that 1 > b > 12> (a− 1). Next, note that we have the following equalities:

    log(1 + az) −az

    1 + z=

    ∫ az0

    (1

    1 + t−

    1

    1 + z

    )dt

    =

    ∫ bz0

    z − t

    (1 + z)(1 + t)dt +

    ∫ zbz

    (1

    1 + t−

    1

    1 + z

    )+

    ∫ azz

    z − t

    (1 + z)(1 + t)dt . (19)

    The second integral in (19) is nonnegative (because its integrand is nonnegative), so

    we complete the proof now by showing that the sum of the first and third integrals

    in (19) is nonnegative:∫ bz0

    z − t(1 + z)(1 + t)

    dt+

    ∫ azz

    z − t(1 + z)(1 + t)

    dt ≥ (1− b)z2b

    (z + 1)2+−(a− 1)2z2

    (z + 1)2≥ 0 ,

    where the two inequalities come from recalling that b < 1, and, respectively, (1− b) =455(a− 1) and b ≥ (a− 1).

    Lemma 6.2 (Lemma 4.4). The function r6(m) =m−1.31

    e2.1m+3.41

    m+1 −1satisfies:

    0.14m > r6(m) > 0.13m− 0.5 , (20)

    where the upper bound holds for m ≥ 2, and the lower bound holds for m ≥ 20.

    Proof. We first prove the upper bound. By the second inequality in (15), (e2.1m+3.41

    m+1 −1) > 0 for m ≥ 2. Thus our desired upper bound in (20) is equivalent to the following:

    m(0.14e2.1m+3.41

    m+1 − 1.14) > − 1.31 ,

    which holds (for positive m) whenever (0.14e2.1m+3.41

    m+1 − 1.14) > 0. This inequality inturn is equivalent to the following (since log is an increasing function):

    2.1m+ 3.41 > (m+ 1) log

    (1.14

    0.14

    )≈ (m+ 1)(2.10) ,

    which is true for positive m, so the proof of the upper bound is complete.

    For the lower bound, by clearing the denominator and gathering exponential terms

    on the right-hand side, we see that the desired inequality is equivalent to the following:

    1.13m− 1.81 ≥ e2.1m+3.41

    m+1 (0.13m− 0.5) .

    We prove this now. The first inequality below is equivalent to the inequality .00004m ≥−2.536, which is true for positive m:

    1.13m− 1.81 ≥ 8.692(0.13m− 0.5)

    > e2.1(20)+3.41

    (20)+1 (0.13m− 0.5) ≥ e2.1m+3.41

    m+1 (0.13m− 0.5) ,

    and the final inequality holds for m ≥ 20 because e2.1m+3.41

    m+1 is a decreasing function

    for positive values of m.460

    22

  • Appendix B: graphs for the proof of Thoeorem 5.1

    Figures 2a–3a below present the graphs of the determinant of the Jacobian matrix

    evaluated at the steady states x∗ and x# (as described in the proof of Theorem 5.1)

    for the network K̃m,n for odd 5 ≤ n ≤ 11 as functions of m. Note that the graphsare nonzero for 2 ≤ m ≤ 5, confirming Conjecture 2.10 for odd 5 ≤ n ≤ 11 and those465values of m. Also, the graphs strongly suggest that the conjecture holds for larger m

    as well. All graphs were made using Desmos Graphing Calculator [13].

    (a) det(df(x∗)) (b) det(df(x#))

    Figure 2: K̃m,5

    (a) det(df(x∗)) (b) det(df(x#))

    Figure 3: K̃m,7

    (a) det(df(x∗)) (b) det(df(x#))

    Figure 4: K̃m,9

    23

  • (a) det(df(x∗)) (b) det(df(x#))

    Figure 5: K̃m,11

    24

    IntroductionBackgroundMass-action kinetics systemsThe sequestration network Km,n

    Constructing multiple steady states via the determinant optimization methodResolving Conjecture 2.10 for the n=3 caseReaction rate constants for which K"0365Km,3 is multistationaryBounding rates and steady states of K"0365Km,3Proving nondegeneracy of steady states for the network K"0365Km,3

    Resolving Conjecture 2.10 for small values of m and nDiscussion


Recommended